You are on page 1of 10

Applied Catalysis A: General 303 (2006) 6271 www.elsevier.

com/locate/apcata

Water-gas shift reaction over Cu/ZnO and Cu/ZnO/Al2O3 catalysts prepared by homogeneous precipitation
Tetsuya Shishido a, Manabu Yamamoto b, Dalin Li b, Yan Tian a, Hiroyuki Morioka c, Masahide Honda c, Tsuneji Sano b, Katsuomi Takehira b,*
a Department of Chemistry, Tokyo Gakugei University, Nukui-Kita 4-1-1, Koganei, Tokyo 184-8501, Japan Department of Chemistry and Chemical Engineering, Graduate School of Engineering, Hiroshima University, Kagamiyama 1-4-1, Higashi-Hiroshima 739-8527, Japan c Hiroshima Prefectural Institute of Industrial Science and Technology, Kagamiyama 3-10-32, Higashi-Hiroshima 739-0046, Japan b

Received 8 November 2005; received in revised form 7 January 2006; accepted 22 January 2006 Available online 13 March 2006

Abstract Both binary Cu/ZnO and ternary Cu/ZnO/Al2O3 catalysts were prepared by homogeneous precipitation (hp) using urea hydrolysis. The structure and the activity for the water-gas shift reaction of these catalysts were studied compared with those prepared by coprecipitation (cp). The binary precursors contained hydroxycarbonates such as malachite and aurichalcite phases, whereas the ternary precursors were composed of hydrotalcite, malachite and aurichalcite phases depending on the metal composition. After thermal decomposition, both catalysts contained apparently CuO and ZnO as crystalline phase. No phase derived from Al was observed, since the amount of Al was small as 10 at.% in the ternary catalysts. After reduction pretreatment with hydrogen, the catalysts were tested for the shift reaction between 150 and 300 8C. The activity of hpcatalysts was higher than that of cp-catalysts; binary hp-Cu/ZnO showed higher activity than ternary hp-Cu/ZnO/Al2O3 catalysts none the less the surface area was larger for the latter than for the former. The activity apparently depended on the surface area of Cu metal formed on the surface of hp-catalysts and a good correlation was observed between the Cu metal particle size and the activation energy of the shift reaction. However, more precise evaluation of the activity based on turn-over frequency strongly suggested the formation of Cu+ species as the active sites at the boundary between Cu metal particles and ZnO particles. Even after the pre-reduction at the high temperature, 250 8C, hp-Cu/ZnO catalyst showed no signicant deactivation as well as no detectable sintering of the Cu metal particles during 50 h of the reaction, indicating that the hp-preparation method afforded the Cu catalysts with high sustainability in the shift reaction. # 2006 Elsevier B.V. All rights reserved.
Keywords: Hydrogen; Water-gas shift reaction; Cu/ZnO catalyst; Homogeneous precipitation; Cu+ species

1. Introduction The water-gas shift reaction is mostly used in the production of hydrogen via the steam reforming of hydrocarbons and is of importance for future energy technologies such as fuel cells. Polymer electrolyte fuel cells (PEFCs), utilizing a protonic conductor, have been extensively studied due to their attractive properties, such as high power density, low emissions (NOx, dust, noise and so on), low temperature operation and compactness [1,2]. In this system, hydrogen is used as a fuel; it is supplied from steam reforming of hydrocarbons such as

* Corresponding author. Tel.: +81 82 424 7744; fax: +81 82 424 7744. E-mail address: takehira@hiroshima-u.ac.jp (K. Takehira). 0926-860X/$ see front matter # 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.apcata.2006.01.031

methane, propane and kerosene. The problem is that the reformed gas contains CO at the level of 110% which adsorbs irreversibly on the Pt electrode of the PEFC at the operating temperature (ca. 80 8C) and hinders the electrochemical reaction [3,4]. Therefore, CO must be removed from the reformed gases to less than 1020 ppm before feeding the gas mixture to the Pt electrode. The shift reaction is desirable for removal of a large amount of CO since it is moderately exothermic reaction (DH298 = 41.1 kJ mol1) and the reaction temperature is easy to control. The equilibrium conversion of CO is dependent largely on the reaction temperature: since the shift reaction is an exothermic reaction, lower temperature is favored for higher CO removal. On the other hand, from the viewpoint of kinetics, the reactant gases are not active enough to reach the chemical equilibrium at low temperature.

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271

63

The application of hydrogen production to PEFCs involves frequent start-ups and shut-downs. The catalysts for the shift reaction are, therefore, sometimes exposed to water and/or oxygen containing atmosphere at low temperatures. Copperbased catalysts are generally more active for the shift reaction, but are more unstable to oxidant gases than precious metals. It is, therefore, important to develop highly stable Cu-based catalysts or highly active precious metal catalysts. Various ceria supported metal catalysts have been prepared and used in the shift reaction; Cu/CeO2 and Ni/CeO2 were prepared by the urea coprecipitationgelation [5], Pd/CeO2 was prepared by impregnation [6], Au/CeO2 was prepared by deposition precipitation [7], and both Au/CeO2 [8] and Cu-Pd/CeO2 catalysts [9] were prepared by the urea coprecipitation gelation method. However, Zalc et al. [10] reported that a large amount of H2 leads to the irreversible reduction of CeO2 and the catalyst deactivation. Pt/TiO2 [11] and Pt/Al2O3 [12] were promising candidate as the catalysts for low-temperature shift reaction. Ru was also active for the shift reaction; Ru/V2O3 catalyst was prepared by the reduction of Ru/V2O5 at 400 8C in H2 [13] and Ru/hydroxyapatite catalyst was prepared by depositionprecipitation method [14]. Ternary Cu/ZnO/Al2O3 catalysts have been widely employed commercially since the early 1960s in the watergas shift reaction; the catalyst was usually prepared by coprecipitation method to afford the higher Cu metal dispersion in the resulting catalyst and, as a consequence, the higher catalytic activity [15]. Coprecipitated Cu/ZnO/Al2O3 catalysts were more sustainable than the impregnated catalysts in the shift reaction [16]. Impregnation of CeO2 on Cu-Zn precursor before calcination was useful for stabilizing the Cu/ZnO catalyst against the shift reaction [17]. High activity was obtained by the formation of spinel type Cu-Al-Zn oxides giving rise to ne Cu particles deposited on the support [18], on which oxygen assisted the shift reaction [19]. The Cu-Al-Zn oxides catalyst is unstable and gave rise to the sintering of Cu at high temperature, whereas stable spinel CuAl2O4 structure was moderately active in the shift reaction. Interestingly, spinel CuMn2O4 catalyst showed high activity as well as high durability in the shift reaction [2022]. When the spinel oxides were calcined above 900 8C, Cu2+ was easily reduced and this may be responsible for the high activity of the spinel compounds. Partial substitution of Fe or Al for Mn enhanced the catalytic activity and, moreover, urea homogeneous co-precipitation method and solgel method, such as citric acid complex method, were effective for the preparation of active Cu-Mn spinel catalyst. We have been elaborating the preparation method of highly active supported Ni catalysts for steam, dry or autothermal reforming of CH4 by citrate method [2325] and coprecipitation method [2629]. Recently, we have reported that Cu/ZnO and Cu/ZnO/Al2O3 catalysts were prepared by homogeneous precipitation (hp) using urea hydrolysis and were successfully applied for hydrogen production by steam reforming of methanol [30]. The catalysts showed higher activities than those prepared by coprecipitation (cp). It was suggested that the good catalytic performances of hp-Cu/ZnO are due to both highly dispersed Cu metal species and to high

accessibility of the Cu metal species to methanol and steam. It was concluded that the homogeneous precipitation method by urea hydrolysis is preferable for the preparation of active Cu/ ZnO catalysts. It is frequently pointed out that reverse water-gas shift reaction signicantly contributes to the product selectivity in the steam reforming of methanol, since the reforming reaction produces actually CO2 at relatively high pressure [31 33]. In the present work, we report the high and stable activity of hp-Cu/ZnO catalysts prepared by the urea homogeneous precipitation method in the shift reaction. 2. Experimental 2.1. Preparation of the catalysts Binary Cu/ZnO and ternary Cu/ZnO/Al2O3 catalysts were prepared by homogeneous precipitation (hp) by urea hydrolysis [3436] as follows: urea was mixed into a solution of corresponding metal nitrates at room temperature, and the urea was hydrolyzed by heating the mixture at 90 8C. During the hydrolysis of urea, hydroxide ions are generated in the homogeneous solution (Eq. (1)) and work as precipitants of the metal nitrates. CONH2 2 H2 O ! 2NH4 HCO3 OH (1)

It is expected that the homogeneity of the precipitate obtained by this method will be higher than that prepared by coprecipitation, since there is no gradient in concentration of precipitants in the solution. The precipitates were then ltered, washed with distilled water, dried at 80 8C and nally calcined at 300 8C for 3 h in air. As reference, coprecipitation (cp) was applied for the preparation of cp-Cu/ZnO and cp-Cu/ZnO/ Al2O3 catalysts; the aqueous solution of corresponding metal nitrates was dropped into aqueous solution of Na2CO3 with vigorous stirring and the pH was adjusted at 7 with NaOH. The obtained precipitate was dried at 80 8C, followed by calcination at 300 8C for 3 h in air. It was conrmed that Cu, Zn and Al were perfectly incorporated in the catalysts by ICP analyses (vide infra), even though Cu2+ ions can be depleted by ammonia originated from urea hydrolysis: the results of ICP analyses of hp-Cu/ZnO (45/55) and hp-Cu/ZnO/Al2O3 (45/45/10) showed the values of 45.5/54.5 (Cu/Zn) and 45.3/ 44.4/10.3 (Cu/Zn/Al) mol%, respectively, where the numbers in the parenthesis showed the metal ratio used in the catalyst preparation. 2.2. Characterization of the catalysts The structure of the catalysts was studied by using XRD, EXAFS, STEM, ICP, TPR, N2O decomposition and N2 adsorption method. Powder X-ray diffraction was recorded on a Rigaku powder diffraction unit, RINT 2250VHF, with mono-chromatized Cu Ka radiation (l = 0.154 nm) at 40 kV and 300 mA. The diffraction pattern was identied by comparing with those included in the JCPDS (Joint Committee of Powder Diffraction

64

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271

Standards) data base. X-ray absorption spectroscopic measurements were performed with synchrotron radiation at a beam line BL7C station of the photon factory, at the High Energy Accelerator Research Organization (Tsukuba, Japan), operated at 2.5 GeV with about 35380 mA of ring current. The data were recorded in transmission mode at room temperature using Si(1 1 1) double crystal monochrometer. Energy was calibrated with Cu K-edge absorption (8981.0 eV), and the energy step of measurement in the XANES region was 0.3 eV. The absorption was normalized to 1.0 at an energy position of 50 eV higher than the absorption edge. Scanning transmission electron microscope measurements were performed with JEOL JEM3000F by using a Noran Voyager energy dispersive X-ray spectroscopy at an accelerating voltage of 300 kV. Inductively coupled plasma (ICP) optical emission spectrometry was used for the determination of the metal content in each sample synthesized above. The measurements were performed with a Perkin-Elmer OPTIMA 3000, and the sample was dissolved in a mixture of HF and HNO3 acids before the measurements. Temperature-programmed reduction (TPR) of the catalyst was performed at a heating rate of 10 8C min1 from ambient temperature to 500 8C using a mixture of 5 vol.% H2/Ar as reducing gas after passing through a 13X molecular sieve trap to remove water. A U-shaped quartz tube reactor, the inner diameter of which was 6 mm, equipped with a TCD for monitoring the H2 consumption was used. Prior to the TPR measurement, 50 mg of sample was calcined at 300 8C for 1.5 h in Ar/O2 (40/10 ml min1) mixed gas ow. The copper metal surface areas were determined by N2O decomposition method at 90 8C as reported by Evans et al. [37], assuming a reaction stoichiometry of two Cu atoms per oxygen atom and a Cu surface density of 1.63 1019 Cu atom m2. Prior to the measurement, 150 mg of sample was reduced at 350 8C for 0.3 h in N2/H2 (30/5 ml min1) mixed gas ow. N2 adsorption (196 8C) study was used to obtain BET surface area of the mixed oxides. The measurement was carried out on a Micromeritics Flow Sorb II2300. The samples after the calcination were pretreated in vacuum at 300 8C for 30 min before the measurements. 2.3. Catalytic reactions and analyses of the products The water-gas shift reaction was carried out using a xed bed reactor at atmospheric pressure. A U-shaped Pyrex glass tube with inner diameter of 4 mm was used as a reactor. Typically 200 mg of catalyst, which has been pelletized and sieved to 0.250.42 mm in diameter, was loaded into the reactor together with 200 mg of quartz beads. The catalyst was treated in a mixed gas ow of N2/H2 (30/5 ml min1) at 350 8C for 20 min and then purged with He (purity, 99.99%). The reaction was started by introducing a gas mixture of 1.45% CO in water and nitrogen to the reactor. Typically mixed gas ow of CO/ H2O/N2 (0.7/2.2/47.1 ml-NTP min1) was supplied for the reaction, where N2 was used as the internal standard for the calculations of the conversions of CO and H2O. H2O was fed by passing a mixed gas of CO and N2 into water in the reservoir which was controlled at the desired temperature.

The products were analyzed by on-line gas chromatographs column (3 m, (GC). A GC with packed molecular sieves 5 A 1/4), He carrier gas and TCD were used to analyze N2 and CO. column (3 m, Another GC with packed molecular sieves 5 A 1/4), Ar carrier gas and TCD were used to analyze H2, N2 and CO. CO2 and H2O were also quantied by another GC with TCD using a packed PEG 20M column (1 m, 1/4). All the lines and valves through the water feed, the reactor, the exit and the gas chromatographs were heated to 130 8C to prevent the condensation of water. 3. Results and discussion 3.1. Crystal structure of the catalysts Fig. 1 shows the XRD patterns of hp-Cu/ZnO catalysts before calcination (A) and after reaction (B). Before calcination, hp-Cu/ZnO samples as deposited showed the patterns of

Fig. 1. XRD patterns of hp-Cu/ZnO catalysts before calcination (A) and after reaction (B). a, hp-Cu; b, hp-Cu/ZnO (90/10); c, hp-Cu/ZnO (80/20); d, hp-Cu/ ZnO (70/30); e, hp-Cu/ZnO (60/40); f, hp-Cu/ZnO (50/50); g, hp-Cu/ZnO (40/ 60); h, hp-Cu/ZnO (30/70); i, hp-Cu/ZnO (20/80); j, hp-Cu/ZnO (10/90). (*), Aurichalcite (Cu,Zn)5(CO3)2(OH)16; (*), malachite Cu2(CO3)(OH)2; (&), ZnO; (&), Cu metal.

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271

65

three types of hydroxycarbonates depending on the metal compositions, i.e. malachite Cu2(CO3)(OH)2 and aurichalcite (Cu,Zn)5(CO3)2(OH)16 (Fig. 1A). All hydroxycarbonates decomposed to CuO and ZnO after calcination at 300 8C for 3 h (data not shown), and were reduced to Cu metal and ZnO after reduction treatment (data not shown). After the shift reaction at 200 8C for 50 h, line broadenings in the reections of Cu metal were observed (Fig. 1B) and these signicantly depended on the preparation method (vide infra). Fig. 2 reveals the XRD patterns of hp-Cu/ZnO/Al2O3 catalysts before calcination (A) and after reaction (B). In the catalysts as deposited, the patterns of four types of hydroxycarbonates were observed; hydrotalcite Cu2Zn4Al2(OH)16CO34H2O and malachite in Cu-rich compositions, and those of aurichalcite and Zn4CO3(OH)6H2O appeared with increasing Zn content (Fig. 2A). Hydrotalcite phase appeared by the addition of Al component into Cu-Zn system. Phase transitions of the ternary systems were almost similar to the binary systems; decomposition to CuO and ZnO, followed by the reduction to Cu.

After the shift reaction at 200 8C for 50 h, the reections of Cu metal and ZnO were still observed and both lines were broader in hp-Cu/ZnO/Al2O3 catalysts than in hp-Cu/ZnO catalysts, suggesting that Al component stabilized ne Cu and ZnO particles. STEM-EDS images of hp-Cu/ZnO/Al2O3(45/ 45/10) catalysts after the shift reaction (data not shown) indicate that Cu and Zn are quite separately distributed each other, while Al is located together with Zn. The calculated results from the line width in XRD showed that the average sizes of Cu metal particles on the catalysts are less than 20 nm, whereas those of ZnO are less than 30 nm. 3.2. XANES and EXAFS spectra of the catalysts Fig. 3 shows the Cu K-edge XANES spectra (A) and the Fourier transforms of k3-weighted Cu K-edge EXAFS spectra (B) of the Cu catalysts before reduction and the reference compounds. Cu foil showed the edge at 8979 eV and two characteristic peaks at the higher energy [37]. Cu2O (Cu+ coordinated by two oxygen atoms) and CuO (Cu2+ in a squareplaner symmetry) showed clearly different XANES spectra. A weak pre-edge ascribed to a dipole-forbidden transition of 1s ! 3d orbital [3740] appears around 8984 eV for the Cu2+ species in CuO. Cu+ compounds such as Cu2O show an intense peak around 8982 eV, attributable to dipole-allowed transition of 1s ! 4p orbital [40,41]. cp-Cu/ZnO (50/50) (Fig. 3A(d)) showed an almost same spectrum as CuO. No signicant difference between cp-Cu/ZnO (50/50) and hp-Cu/ZnO (50/50) (Fig. 3A(e)) was observed in XANES spectra. In the Fourier transforms for both cp-Cu/ZnO (50/50) and hp-Cu/ZnO (50/50) (Fig. 3B), peaks due to Cu-O and Cu-O-Cu were clearly (non-phase-shift corrected), observed at 1.5 and 2.5 A respectively. The position of these peaks is quite similar to CuO. The comparison of these spectra with those for the reference suggests that the copper on both samples exists as the CuO species with Cu2+ in a distorted octahedral coordination. The peak intensity at 8984 eV for both hp-Cu/ZnO (50/50) and hp-Cu/ZnO/Al2O3 (45/45/10) was slightly weaker than those for CuO and cp-Cu/ZnO (50/50), suggesting that the Cu2+ in the hp-catalysts possesses the octahedral coordination less distorted than those in CuO and the cp-catalyst. Moreover, the edge positions slightly shifted toward the higher energy for the hp-catalysts compared with those for CuO and the cp-catalyst, suggesting that the electron density around Cu2+ in the former increased compared to that around Cu2+ in the latter. In the EXAFS spectra (data not shown), the phases of the EXAFS oscillation for cp-Cu/ZnO (50/50), hp-Cu/ZnO (50/50) and hpCu/ZnO/Al2O3 (45/45/10) were quite similar, whereas the amplitude for hp-Cu/ZnO (50/50) and hp-Cu/ZnO/Al2O3 (45/ 45/10) were weaker than that for cp-Cu/ZnO (50/50). This indicates that the particle size of CuO is smaller in the hpcatalysts than in the cp-catalyst, coinciding with the results observed by XRD analyses in the previous paper [30]. Among two hp-catalysts, the addition of Al caused a further decrease in the EXAFS oscillation; smaller-sized CuO particles were formed in hp-Cu/ZnO/Al2O3 (45/45/10), also coinciding with the results obtained by XRD analyses (vide supra).

Fig. 2. XRD patterns of hp-Cu/ZnO/Al2O3 catalysts before calcination (A) and after reaction (B). a, hp-Cu/ZnO/Al2O3 (80/10/10); b, hp-Cu/ZnO/Al2O3 (70/20/ 10); c, hp-Cu/ZnO/Al2O3 (60/30/10); d, hp-Cu/ZnO/Al2O3 (50/40/10); e, hp-Cu/ ZnO/Al2O3 (40/50/10); f, hp-Cu/ZnO/Al2O3 (30/60/10); g, hp-Cu/ZnO/Al2O3 (20/70/10); h, hp-Cu/ZnO/Al2O3 (10/80/10). (*), Aurichalcite (Cu,Zn)5(CO3)2(OH)16; (*), malachite Cu2(CO3)(OH)2; (~), Zn4CO3(OH)6H2O; (~), hydrotalcite Cu2Zn4Al2(OH)16CO34H2O; (&), ZnO; (&), Cu metal.

66

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271

Fig. 3. Cu K-edge XANES spectra (A) and Fourier transforms of k3-weighted Cu K-edge EXAFS spectra (B) of the Cu samples. (a) Cu foil, (b) Cu2O, (c) CuO, (d) cp-Cu/ZnO (50/50), (e) hp-Cu/ZnO (50/50), (f) hp-Cu/ZnO/Al2O3 (45/45/10).

3.3. TPR of the catalysts The TPR results of both hp-Cu/ZnO and hp-Cu/ZnO/ Al2O3 catalysts are shown in Fig. 4. Although the reduction of bulk ZnO is thermodynamically feasible at high temperatures, ZnO showed no distinct peak of the reduction up to 350 8C (Fig. 4A(k)). CuO showed a gradual reduction as a shoulder around 200 8C followed by a clear reduction peak around 240 8C (Fig. 4A(a)). Starting from ZnO-rich samples, a weak peak appeared around 220 8C by the addition of Cu, and further a new peak appeared around 240 8C when the Cu/ Zn ratio exceeded 50/50 (Fig. 4A(f)). Also in the TPR of hpCu/ZnO/Al2O3 samples (Fig. 4B), a weak peak was observed around 220 8C at the Cu content below 30 % (Figs. 4B(f and g)). The peak around 240 8C appeared at the Cu content of 20 % (Fig. 4B(g)) and a further increase in the Cu content quickly resulted in a unication of the two peaks to a single broad peak. This strongly suggests that Al2O3 homogenized and stabilized both components in the Cu-Zn binary system. Fierro and co-workers [42] reported that Cu/ZnO/Al2O3 calicned at 400 8C for 4 h showed only one reduction peak, centered at 223 8C, which was assigned to the reduction of CuO to Cu: CuO + H2 ! Cu0 + H2O. The present samples were calcined under the milder conditions, i.e. at 300 8C for 4 h, and therefore the effects of the precursors formed may be observed more evidently. Although ZnO was not reduced under our experimental conditions, partial reduction of surface ZnO, which may lead to the formation of a-brass (i.e. a dilute alloy of zinc in copper) only few layers thick on the copper crystallites during the catalyst reduction, cannot be ruled out. In fact, detailed thermodynamic calculations [43] are in favor of such a hypothesis, showing that the equilibrium zinc content in a surface a-brass is about 5% during catalyst reduction at 300 8C. The shapes of the reduction peaks (Fig. 4A(a and b) and B(ae)), which are asymmetric with a tail towards lower temperature, may be the result of overlapping of several elemental reduction processes arising from different Cu2+ species. The peak of CuO reduction in both Cu/ZnO and Cu/

Fig. 4. TPR proles of hp-Cu/ZnO (A) and hp-Cu/ZnO/Al2O3 (B) catalysts. (A) a, hp-Cu; b, hp-Cu/ZnO (90/10); c, hp-Cu/ZnO (80/20); d, hp-Cu/ZnO (70/30); e, hp-Cu/ZnO (60/40); f, hp-Cu/ZnO (50/50); g, hp-Cu/ZnO (40/60); h, hp-Cu/ ZnO (30/70); i, hp-Cu/ZnO (20/80); j, hp-Cu/ZnO (10/90); k, hp-ZnO. (B) a, hpCu/ZnO/Al2O3 (80/10/10); b, hp-Cu/ZnO/Al2O3 (70/20/10); c, hp-Cu/ZnO/ Al2O3 (60/30/10); d, hp-Cu/ZnO/Al2O3 (50/40/10); e, hp-Cu/ZnO/Al2O3 (40/ 50/10); f, hp-Cu/ZnO/Al2O3 (30/60/10); g, hp-Cu/ZnO/Al2O3 (20/70/10); h, hpCu/ZnO/Al2O3 (10/80/10).

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271

67

ZnO/Al2O3 (Figs. 4A and B) was displaced by more than 50 8C towards lower temperatures with respect to that of CuO at 271 8C [42]. For ternary CuO/ZnO/Al2O3 system, the reduction temperature of copper species on the present catalyst prepared by hp-method almost coincided with that observed for the catalyst prepared by coprecipitation. The TPR proles of the catalyst prepared by coprecipitation showed a single peak with maximum at 223 8C [42]. Binary CuO/ZnO system showed more signicant decrease in the Cu2+ reduction temperature than ternary CuO/ZnO/Al2O3 system (Fig. 4A and B), and moreover on both systems an increase in the Zn content caused a decrease in the Cu2+ reduction temperature (Fig. 4A(i and j) and B(f and g)). In the XRD analyses, these catalysts showed reection line of neither Cu metal (Figs. 1B(i and j) and 2B(f and g)) nor CuO after the calcination (data not shown). These results clearly indicate that ZnO enhances the reducibility of the copper phase, which is probably due to the formation of highly dispersed CuO particles on ZnO. It was reported that ZnO promotes the copper reducibility in the overall range of composition of CuO/ZnO binary system [44]. A single reduction process occurs for catalysts with higher Cu loading, whereas the lower copper content catalysts are characterized by two reduction processes related to two different copper species. One is similar to the CuO species present in the higher copper loading catalyst and the other is represented by well dispersed CuO species in contact with the surface of ZnO particles which are reduced at lower temperatures. This may be correlated with the fact that Cu+ species were observed by Auger-electron spectra (vide infra). 3.4. Activity of the catalysts Activities of hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 catalysts for the shift reaction are shown in Fig. 5A and B, respectively, together with the BET surface area and Cu metal surface area. The marks with asterisk (*) show the reaction results and the surface areas obtained over cp-Cu/ZnO (Fig. 5A) and cp-Cu/ ZnO/Al2O3 (Fig. 5B) catalysts. The activity of hp-Cu/ZnO catalyst was higher than that of cp-Cu/ZnO, none the less the surface area was larger on the latter than on the former. Also the activity of hp-Cu/ZnO/Al2O3 catalyst was higher than that of cp-Cu/ZnO/Al2O3 catalyst. Among hp-Cu/ZnO catalysts, CuO alone showed a weak activity for the shift reaction, and the binary system formation with ZnO resulted in a clear enhancement in the activity (Fig. 5A). Although the BET surface area increased, more signicant enhancement was observed in the surface area of metallic copper by the binary system formation for Cu/ZnO catalysts. Also for the ternary system, i.e. hp-Cu/ZnO/Al2O3 catalysts, the activity of Cu was enhanced by coupling with ZnO. In the presence of Al component, the surface area monotonously decreased with increasing Zn component, whereas the Cu metal surface area showed almost same behavior as the Cu/ZnO binary system, i.e. the highest value was obtained at 50% of Zn component and decreased both with increasing and with decreasing Zn component. Although the surface areas of hp-Cu/ZnO/Al2O3 catalysts are larger than those of hp-Cu/ZnO catalysts, the CO
Fig. 5. Activity and physical properties of hp-Cu/ZnO (A) and hp-Cu/ZnO/ Al2O3. (B) catalysts. Catalyst, 200 mg; ow rate: CO/H2O/N2 = 0.7/2.2/47.1 mlNTP min1; reaction temperature, 150 8C. (*), CO conversion; (~), Cu0 surface area; (&), catalyst surface area. The marks with asterisk show the results obtained over cp-Cu/ZnO (A) and cp-Cu/ZnO/Al2O3 (B) catalysts.

conversion was higher on the latter than on the former. It is likely that the activities of both hp-Cu/ZnO and hp-Cu/ZnO/ Al2O3 catalysts substantially depend on the Cu metal surface area than on the BET surface area. 3.5. Activation energy and turn over frequency of the catalysts Apparent activation energies in the shift reactions and the particle sizes of copper metal after the reaction are shown in Fig. 6 for hp-Cu/ZnO (A) and hp-Cu/ZnO/Al2O3 (B) catalysts. Both activation energy and Cu metal particle size increased with increasing Cu content, especially the increase was signicant at the Cu content above 6070%. Activation energies were lower over hp-Cu/ZnO than over hp-Cu/ZnO/ Al2O3 catalysts with the Cu content below 80 at.%, whereas the Cu metal particle sizes were smaller on hp-Cu/ZnO/Al2O3 than on hp-Cu/ZnO catalysts, coinciding with the results obtained by XRD. Turn over frequencies (TOF) over hp-Cu/ZnO and hp-Cu/ ZnO/Al2O3 catalysts were obtained for the shift reaction at the reaction temperature of 110 8C and at GHSV = 60,000 1 ml h1 g cat (Fig. 7). When TOF values were calculated based

68

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271

Fig. 7. Turn over frequencies of hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 catalysts. (*) and (*), hp-Cu/ZnO; (&) and (&), hp-Cu/ZnO/Al2O3; ( ), calculated based on Cu0 obtained by N2O pulse method; (), calculated based on Cu+ obtained from Auger-electron spectra.

Fig. 6. Activation energy and Cu metal particle size of hp-Cu/ZnO (A) and hpCu/ZnO/Al2O3 (B) catalysts. (*), Cu metal particle size; (&), apparent activation energy.

on Cu metal surface area, i.e. assuming two Cu atoms per oxygen atom and a Cu surface density of 1.63 1019 Cu atom m2 in N2O decomposition (dotted lines), the values decreased with increasing Cu content in both hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 catalysts. This well agreed with the clear dependency of activation energy on Cu content (Fig. 6), suggesting that the nature of the Cu active site varies depending on the Cu content on both hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 catalysts. Moreover, hp-Cu/ZnO showed the higher TOF values than hp-Cu/ZnO/Al2O3, contrarily to the results obtained previously in the steam reforming of methanol [30]. 3.6. Active sites on hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 The precursors of hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 catalysts were prepared by urea method, which allows the preparation of different hydrotalcite-like compounds containing Mg-Al, or Zn-Al, Ni-Al and Mg-Fe having a high crystalline degree and a narrow particle-size distribution [45]. The presence of Cu2+ cations makes the synthetic procedure more complex, at least for two reasons [46]. First, Cu2+ ions show the JahnTeller effect that favors the formation of distorted octahedral structures and preferentially gives rise to the precipitation of malachite phases. Secondarily, Cu2+ ions can be depleted by ammonia originated from urea hydrolysis. Actually in the present work, various phases, i.e. hydrotalcite, malachite, aurichalcite and Zn4CO3(OH)6H2O, were observed

during the preparation of hp-Cu/ZnO/Al2O3, whereas malachite and aurichalcite were for hp-Cu/ZnO. It is likely that these phases substantially affected the structure after the calcination and nally followed by the reduction. Auger-electron spectra of hp-Cu/ZnO and hp-Cu/ZnO/ Al2O3 catalysts after the reduction are shown in Fig. 8A and B, respectively. According to the results reported by Batista et al. [47], Cu0 and Cu+ species have the following values of the kinetic energy (Cu-LMM): 918.6 and 916.5 eV, respectively. A peak assigned to Cu+ was clearly observed at 917.0 eV together with that of Cu0 at 919.0 eV in the Cu-LMM Auger spectrum of both hp-Cu/ZnO (Fig. 8A) and hp-Cu/ZnO/Al2O3 (Fig. 8B) varying their intensities dependently on the Cu/Zn ratio. Moreover a satellite peak characteristic of Cu2+ was unobservable around 945 eV in the XPS of Cu 2p3/2 on both hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 (data not shown) [48]; thus, the oxidation state of Cu could be zero- or mono-valent. It is therefore concluded that Cu+ species was formed on both hp-Cu/ZnO and hp-Cu/ZnO/Al2O3 after the reduction. Auger peaks were more intensive for hp-Cu/ZnO than for hp-Cu/ ZnO/Al2O3. The peak area ratio of Cu+/Cu0 obviously increased, reached the maximum value and decreased with increasing Cu content for hp-Cu/ZnO, whereas the ratio increased and showed almost constant value with increasing Cu content for hp-Cu/ZnO/Al2O3. These phenomena well coincided with those observed for the activity shown as CO conversion (Fig. 5), suggesting that the Cu+ species works as the active site for the shift reaction. The values of TOF were re-calculated based on the surface amount of Cu+ estimated from Auger peak ratio of Cu+/Cu0 assuming that Cu+ species was localized on the catalyst surface (Fig. 7). The TOFs converged to constant values independently on the Cu content on both hp-Cu/ZnO (0.67 s1) and hp-Cu/ZnO/Al2O3 (0.58 s1), strongly supporting the hypothesis of the Cu+ species as the active site.

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271

69

Fig. 8. Auger electron spectra of hp-Cu/ZnO (A) and hp-Cu/ZnO/Al2O3 (B) after reduction. (A) a, hp-Cu/ZnO (90/10); b, hp-Cu/ZnO (80/20); c, hp-Cu/ZnO (70/30); d, hp-Cu/ZnO (60/40); e, hp-Cu/ZnO (50/50); f, hp-Cu/ZnO (40/60); g, hp-Cu/ZnO (30/70); h, hp-Cu/ZnO (20/80); i, hp-Cu/ZnO (10/90). (B) a, hp-Cu/ZnO/Al2O3 (80/10/10); b, hp-Cu/ZnO/Al2O3 (70/20/10); c, hp-Cu/ZnO/Al2O3 (60/30/10); d, hp-Cu/ZnO/Al2O3 (50/40/10); e, hp-Cu/ZnO/Al2O3 (45/45/10); f, hp-Cu/ZnO/Al2O3 (40/50/10); g, hp-Cu/ZnO/Al2O3 (30/60/10); h, hp-Cu/ZnO/Al2O3 (20/70/10); i, hp-Cu/ZnO/Al2O3 (10/80/10). All samples were pre-reduced in a mixed gas ow of N2/H2 (30/5 ml min1).

There are various theories describing the nature of Cu-ZnO interactions in the catalyst [49]. For instance, some authors have suggested that Cu is incorporated in the ZnO phase on interstitial and substitutional sites, assuming three possible valence states (Cu0, Cu+ and Cu2+) [50,51]. XPS measurements revealed the existence of Cu+ species on the surface of binary Cu/ZnO catalysts prepared via hydroxycarbonate precursors in aqueous solution after exposure to an O2/ CH3OH mixture [52]. It was reported that ZnO segregates after reduction of Cu/ZnO/SiO2 catalyst; the migration of ZnO on top of Cu was followed by the formation of a (partly) oxidized Cu in a Cu+/ZnO surface with oxygen vacancies which worked as the active sites for methanol synthesis [53]. It is likely that the shift reaction was catalyzed via a reductionoxidation mechanism between Cu0 $ Cu+, in which Cu+ sites chemisorbed and oxidized CO to CO2 to form Cu0, whereas the reduced Cu0 sites were reoxidized by H2O to form Cu+ and H2 [54,55]. If the rate determining step would be the rst CO oxidation reaction, the constant TOFs values based on Cu+ could be well explained. None the less of this hypothesis, the important role Cu0 cannot be denied since the reaction proceeded by the reductionoxidation mechanism between Cu0 and Cu+. Moreover, the amount of Cu+ sites reasonably depends on the particle size of Cu metal if Cu+ forms at the boundary between Cu metal particles and ZnO particles, suggesting an apparent dependency of the activity on Cu0 species. As a result, the amount of Cu+ sites increased with decreasing the Cu metal particle size, resulting in an increase in the activity as observed in a decrease in Ea (Fig. 6). 3.7. Effect of the reduction conditions hp-Cu/ZnO catalysts were reduced under the varying conditions, i.e. at varying temperature and for varying time,

and were used in the shift reactions. The activation energies obtained in these reactions are plotted against the particle sizes of Cu metal produced after the reduction treatments (Fig. 9). The reduction at the higher temperature brought the formation of the larger sized Cu metal particles as well as the larger activation energy. Moreover, the activation energy showed a good correlation with the particle size of Cu metal. Interestingly the reduction of hp-Cu/ZnO at 250 8C for 2 h resulted in the lower activation energy than that at 200 8C for 2 h, which is exceptional from the view point of the effect of the reduction temperature on the activity of Cu catalyst. This activation energy was the lowest among those obtained in a series of experiments in Fig. 9. Both hp-Cu/ZnO and the commercial Cu/ZnO/Al2O3 catalysts were reduced at 200 and 250 8C and the effect of the reduction temperature on the sustainability of the Cu catalysts was tested for 50 h (Fig. 10). The commercial catalyst showed the more stable activity when the catalyst was prereduced at the lower temperature of 200 8C. Oppositely, hpCu/ZnO showed the more stable activity by the reduction at the higher temperature of 250 8C, and moreover the activity of this catalyst was higher than that of the commercial Cu/ZnO/Al2O3 catalyst. Binary hp-CuO/ZnO system showed more signicant decrease in the Cu2+ reduction temperature than ternary hpCuO/ZnO/Al2O3 system (Fig. 4A and B), suggesting an easier occurring of the reductionoxidation cycle between Cu0 $ Cu+ on the former than the latter system. Indeed the absence of Al2O3 clearly afforded better catalytic performance for hp-Cu/ZnO catalysts (Figs. 6 and 7), whereas the presence of small amount of alkaline-earth metal oxide such as MgO enhanced the catalytic activity of hp-Cu/ZnO [56]. Both Al2O3 and MgO distributed homogeneously in hp-Cu/ZnO catalyst particles by SEM-EDS observation. It is suggested that Al2O3 deplete the activity of hp-Cu/ZnO catalyst for the shift

70

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271 Table 1 Cu0 particle size on the Cu catalysts before and after the reactiona Catalyst Cu0 particle size (nm) Before Cu/ZnO (45/55) Commercial Cub Cu/ZnO (45/55)c Commercial Cuc
b

After 13.6 13.1 10.1 14.9

9.1 12.0 9.3 12.5

a Reaction time: 50 h; Cu0 particle size was calculated from the results of XRD. b Pre-reduced at 200 8C. c Pre-reduced at 250 8C.

Fig. 9. Activation energy and Cu metal particle size of hp-Cu/ZnO (45/55) catalyst reduced under various conditions. All samples were pre-reduced in a mixed gas ow of N2/H2 (30/5 ml min1) and used in the shift reaction.

indicate that hp-preparation can afford the sustainability on the Cu catalyst even in the absence of Al2O3. This is probably due to the homogeneity of metal components in the catalyst prepared, which results in strong metal-support interactions on the catalyst. 4. Conclusions The structure and the activity of Cu/ZnO and Cu/ZnO/ Al2O3 catalysts prepared by homogeneous precipitation have been studied for the water gas-shift reaction. The catalysts contained apparently CuO and ZnO as crystalline phase in both catalysts; the amount of Al was small as 10 at.% in the ternary catalysts. Homogeneous precipitation afforded high activity as well as high sustainability to both Cu/ZnO and Cu/ ZnO/Al2O3 catalysts, among which the binary catalysts showed higher activity than the ternary catalysts. Interestingly the former showed higher activity than the latter, even though the surface area was larger on the latter than on the former. The activity apparently depended on the Cu metal surface area of hp-catalysts and a good correlation was observed between the Cu metal particle size and the activation energy of the shift reaction. However, turn-over frequency varied signicantly depending on the preparation method as well as the contents of metal components. It was strongly suggested that Cu+ species formed at the boundary between Cu metal particles and ZnO particles and worked as the real active sites according to the more precise evaluation of the activity based on turn-over frequency. hp-Cu/ZnO catalyst pre-reduced at the higher temperature, 250 8C, showed no signicant deactivation as well as no detectable sintering of the Cu metal particles during 50 h of the reaction. Homogeneous precipitation afforded highly dispersed Cu metal particles as well as strong metal-support interaction on the catalysts, resulting in the high and stable activity. References
[1] G.G. Scherer, Solid State Ionics 94 (1997) 249. [2] S. Ahmed, M. Krumpelt, Int. J. Hydrogen Energy 26 (2001) 291. cheerhoff, B. Ho hlein, R. Menzer, U. Stimming, J. [3] V.M. Schmidt, P. Bro Power Sources 49 (1994) 299. [4] F. Vidal, B. Busson, C. Six, O. Pluchery, A. Tadjeddine, Surf. Sci. 502503 (2002) 485. [5] Y. Li, Q. Fu, M. Flytzani-Stephanopoulos, Appl. Catal. B 27 (2000) 179.

reaction, unexpectedly from the results obtained in steam reforming of methanol [30]. A possible explanation of the low activity of hp-Cu/ZnO/Al2O3 catalyst is that Al2O3 suppress the formation of Cu+ species as seen in Fig. 8. On the contrary, a small amount of MgO enhanced the formation of Cu+ species on hp-Cu/ZnO catalyst [56]. It is well known that the commercial Cu catalysts are very sensitive against the reduction treatments and the reduction under a severe condition frequently resulted in the deactivation of the catalysts. This can be well explained by the sintering of Cu metal particles. Actually, after the reaction for 50 h, the average size of Cu metal particles increased over hp-Cu/ZnO pre-reduced at 200 8C, while that showed no signicant change over hp-Cu/ZnO pre-reduced at 250 8C (Table 1). On the contrary, over the commercial Cu catalysts, pre-reduction at 200 8C was rather effective for keeping the Cu metal particle size than that at 250 8C. These results clearly

Fig. 10. Sustainability of hp-Cu/ZnO (45/55) in the shift reaction. hp-Cu/ZnO (45/55) catalysts: (*), reduced at 250 8C; (*), reduced at 200 8C. Commercial Cu/ZnO/Al2O3 catalyst: (~), reduced at 250 8C; (~), reduced at 200 8C.

T. Shishido et al. / Applied Catalysis A: General 303 (2006) 6271 [6] S. Hilaire, X. Wang, T. Luo, R.J. Gorte, J. Wagner, Appl. Catal. A 215 (2001) 271. [7] T. Tabakova, F. Boccuzzi, M. Manzoli, D. Andreeva, Appl. Catal. A 252 (2003) 385. [8] Q. Fu, W. Deng, H. Saltsburg, M. Flytzani-Stephanopoulos, Appl. Catal. B 56 (2005) 57. [9] E.S. Bickford, S. Velu, C. Song, Catal. Today 99 (2005) 347. [10] J.M. Zalc, V. Sokolovskii, D.G. Lofer, Appl. Catal. A 215 (2001) 271. [11] P. Panagiotopoulou, D.I. Kondarides, J. Catal. 225 (2004) 327. [12] T. Utaka, T. Takeguchi, R. Kikuchi, K. Eguchi, Appl. Catal. 246 (2003) 117. [13] T. Utaka, T. Okanishi, T. Takeguchi, R. Kikuchi, K. Eguchi, Appl. Catal. A 245 (2003) 343. [14] A. Venugopal, M.S. Scurrell, Appl. Catal. A 245 (2003) 137. s, N. Amadeo, M. Laborde, C.R. Apestegu a, Appl. Catal. A [15] M.J.L. Gine 131 (1995) 283. [16] Y. Tanaka, T. Utaka, R. Kikuchi, K. Sasaki, K. Eguchi, Appl. Catal. A 238 (2003) 11. [17] M. Rnning, F. Huber, H. Meland, H. Venvik, D. Chen, A. Holmen, Catal. Today 100 (2005) 249. [18] K. Sekizawa, S. Yano, K. Eguchi, H. Arai, Appl. Catal. A 169 (1998) 291. [19] T. Utaka, K. Sekizawa, K. Eguchi, Appl. Catal. A 194195 (2000) 21. [20] Y. Tanaka, T. Utaka, R. Kikuchi, K. Sasaki, K. Eguchi, Appl. Catal. A 242 (2003) 287. [21] Y. Tanaka, T. Utaka, R. Kikuchi, T. Takeguchi, K. Sasaki, K. Eguchi, J. Catal. 215 (2003) 271. [22] Y. Tanaka, T. Takeguchi, R. Kikuchi, K. Eguchi, Appl. Catal. A 279 (2005) 59. [23] T. Hayakawa, H. Harihara, A.G. Andersen, A.P.E. York, K. Suzuki, H. Yasuda, K. Takehira, Angew. Chem. Int. Ed. Engl. 35 (1996) 192. [24] R. Shiozaki, A.G. Andersen, T. Hayakawa, S. Hamakawa, K. Suzuki, M. Shimizu, K. Takehira, J. Chem. Soc. Faraday Trans. 93 (1997) 3235. [25] K. Takehira, T. Shishido, M. Kondo, J. Catal. 207 (2002) 307. [26] A.I. Tsyganok, T. Tsunoda, S. Hamakawa, K. Suzuki, K. Takehira, T. Hayakawa, J. Catal. 213 (2003) 191. [27] K. Takehira, T. Shishido, P. Wang, T. Kosaka, K. Takaki, Phys. Chem. Chem. Phys. 5 (2003) 3801. [28] K. Takehira, T. Shishido, P. Wang, T. Kosaka, K. Takaki, J. Catal. 221 (2004) 43. [29] K. Takehira, T. Kawabata, T. Shishido, K. Murakami, T. Ohoi, D. Shoro, M. Honda, K. Takaki, J. Catal. 231 (2005) 92. [30] T. Shishido, Y. Yamamoto, H. Morioka, K. Takaki, K. Takehira, Appl. Catal. A 263 (2004) 249.

71

gl, R. Schoma cker, Appl. [31] H. Purnama, T. Ressler, R.E. Jentoft, R. Schlo Catal. A 259 (2004) 83. [32] J.K. Lee, J.B. Ko, D.H. Kim, Appl. Catal. A 278 (2004) 25. [33] A. Mastalir, B. Frank, A. Szizybalski, H. Soerijanto, A. Deshpande, M. cker, R. Schlo gl, T. Ressler, J. Catal. 230 (2005) Niederberger, R. Schoma 464. [34] R.J. Candal, A.E. Regazzoni, M.A. Blesa, J. Mater. Chem. 2 (1992) 657. [35] G.J. de, A.A. Soler-Illia, R.J. Candal, A.E. Regazzoni, M.A. Blesa, Chem. Mater. 9 (1997) 184. [36] H. Morioka, H. Tagaya, K. Karasu, J. Kadokawa, K. Chiba, J. Solid State Chem. 117 (1995) 337. [37] S. Velu, K. Suzuki, C.S. Gopinath, H. Yoshida, T. Hattori, Phys. Chem. Chem. Phys. 4 (2002) 1990. [38] G. Meitzner, E. Iglesia, Catal. Today 53 (1999) 433. [39] C. Lamberti, S. Bordiga, G. Spoto, A. Zecchina, F. Geobaldo, G. Vlaic, M. Bellatreccia, J. Phys. Chem. 101 (1997) 344. [40] Y.-J. Huang, H.P. Wang, J.-F. Lee, Appl. Catal. B 40 (2003) 111. [41] Y.-J. Huang, H.P. Wang, J.-F. Lee, Chemosphere 50 (2003) 1035. n-Cabrera, M. Lo pez Granados, J.L.G. Fierro, J. Catal. 210 (2002) [42] I. Melia 273. [43] M.S. Spencer, Surf. Sci. 192 (1987), 323, 336, 329. [44] G. Fierro, M. Lo Jacono, M. Inversi, P. Porta, F. Cioci, R. Lavecchina, Appl Catal. A 137 (1996) 327. [45] U. Costantino, F. Marmottini, M. Nochetti, R. Vivani, Eur. J. Inorg. Chem. (1998) 1439. [46] M. Turco, G. Bagnasco, U. Constantino, F. Marmottini, T. Montanari, G. Raims, G. Busca, J. Catal. 228 (2004) 43. [47] J. Batista, A. Pintar, D. Mandrino, M. Jenko, V. Martin, Appl. Catal. A 206 (2001) 113. [48] Y. Tanaka, R. Kikuchi, T. Takeguchi, K. Eguchi, Appl. Catal. B 57 (2005) 211. nter, T. Ressler, B. Bems, C. Bu scher, T. Genger, O. Hinrichsen, [49] M.M. Gu gl, Catal. Lett. 71 (2001) 37. M. Muhler, R. Schlo [50] K. Klier, Adv. Catal. 31 (1982) 243. [51] V. Ponec, Surf. Sci. 272 (1992) 111. n-Cabrera, J.L.G. Fierro, Appl. Catal. A [52] J. Agrell, M. Boutonnet, I. Melia 253 (2003) 201. [53] W.P.A. Jansen, J. Beckers, J.C.v.d. Heuvel, A.W. Denier v.d. Gon, A. Bliek, H.H. Brongersma, J. Catal. 210 (2002) 229. [54] Y. Li, Q. Fu, M. Flytzani-Stephanopoulos, Appl. Catal. B 27 (2000) 179. [55] N.A. Koryabkina, A.A. Phatak, W.F. Ruettinger, R.J. Farrauto, F.H. Ribeiro, J. Catal. 217 (2003) 233. [56] T. Shishido, M. Yamamoto, I. Atake, Y. Tian, H. Morioka, M. Honda, K. Takaki, T Sano, K. Takehira, J. Mol. Catal. A, submitted for publication.

You might also like